Next Article in Journal
Synthesis and Theoretical Studies of Aromatic Azaborines
Next Article in Special Issue
Divergent Synthesis of Four Monomeric Ellagitannins toward the Total Synthesis of an Oligomeric Ellagitannin, Nobotanin K
Previous Article in Journal
Mo2C as Pre-Catalyst for the C-H Allylic Oxygenation of Alkenes and Terpenoids in the Presence of H2O2
Previous Article in Special Issue
Recent Applications of Pd-Catalyzed Suzuki–Miyaura and Buchwald–Hartwig Couplings in Pharmaceutical Process Chemistry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enantioselective Total Synthesis of Multifidene, a Sex Pheromone of Brown Algae

1
Division of Environmental Materials Science, Graduate School of Environmental Science, Hokkaido University, Sapporo 060-0810, Japan
2
Shimoda Marine Research Center, University of Tsukuba, Shizuoka 415-0025, Japan
*
Author to whom correspondence should be addressed.
Organics 2022, 3(3), 187-195; https://doi.org/10.3390/org3030015
Submission received: 20 April 2022 / Revised: 16 June 2022 / Accepted: 30 June 2022 / Published: 8 July 2022
(This article belongs to the Special Issue New Reactions and Strategies for Natural Product Synthesis)

Abstract

:
The total synthesis of multifidene, a sex pheromone found in brown algae, is described. The synthesis features the highly enantioselective and diastereoselective addition reaction of an aldehyde to a nitroolefin in the presence of a Hayashi–Jørgensen catalyst and a Nef reaction initiated by visible light irradiation. These key reactions enabled the 11-step synthesis from commercially available compounds. The synthetic pheromones are examined with gametes.

1. Introduction

The chemical structure of multifidene (1), obtained from marine brown algae Cutleria multifida as one of its sex pheromones, was first determined by Jaenicke, Müller and Moore in 1974 (Figure 1) [1] as a C11 hydrocarbon. Other C11 hydrocarbon pheromones, such as ectocarpene (2), hormosirene (3) and finavarrene (4), from brown algae have been previously reported [2,3,4,5]. To date, 11 parent pheromones and more than 50 of their stereoisomers [6,7,8,9,10] have been identified in brown algae and several microalgae [11,12,13]. In the brown algae, the C11 hydrocarbons are presumed to have the following functions. In sexual reproduction, C11 hydrocarbons from female gametes contribute to their efficient mating by inducing the release of male gametes and by attracting male gametes [10]. In vegetative thalli, C11 hydrocarbons prevent feeding as a chemical defense [14,15].
In the 1980s, extensive chemical and biological studies of the brown algae through biosynthesis, pheromone-mediated spermatozoid release, chemotaxis and synthesis of the pheromones were reported [7]. Concerning the synthesis of 1, several reports have been published [16,17,18,19,20,21,22,23]. Some of them are shown in Figure 2. In the synthesis of Hemamalini and co-workers, starting from optically active carboxylic acid, a cobalt-mediated radical cyclization reaction was employed to obtain lactone. This lactone was successfully derived to 1 in four steps. Furstoss and co-workers utilized a kinetic resolution of racemic cyclobutanone with the fungus Cnninghamella echinulata to obtain optically active lactone in >98% ee, which was converted to 1. In 2019, our research group revealed the novel function (i.e., control of the phototactic sign) of sex pheromones in the phototaxis of male gametes in the brown alga Mutimo cylindricus with synthetic ectocarpene (2) [23]. Furthermore, we revealed that this conversion of the phototactic sign coincided with the dynamics of the intracellular cyclic nucleotide and Ca2+ concentration. The C11 sex pheromones are biologically synthesized from unsaturated fatty acid 5 [24], as shown in Figure 3. Hydrogen abstraction from an active methylene followed by single-electron oxidation generates a carbocation, which is further subjected to a decarboxylation reaction, giving various intermediates and/or final compounds. As a result, the C11 pheromones are released from female gametes as a mixture. It is possible that each pheromone may show different functions against gametes. In this paper, we describe the enantioselective total synthesis of multifidene (1) through a conjugate addition reaction by a Hayashi–Jørgensen catalyst and the Nef reaction under visible right irradiation as key steps in order to study the function of each C11 pheromone and develop molecular probes to identify the mechanisms of action.

2. Materials and Methods

2.1. General Methods

The IR spectra were recorded on a JASCO FTIR-4100 Type A spectrometer (JASCO corporation, Tokyo, Japan) using a NaCl cell. The 1H NMR and 13C NMR spectra were recorded using a JNM-EX 400 (400 MHz and 100 MHz) spectrometer (JEOL Ltd., Tokyo, Japan). Chemical shifts were reported in ppm relative to CHCl3 in CDCl3 for 1H NMR (δ = 7.26) and 13C NMR (δ = 77.0). Splitting patterns for 1H NMR were designated as “s, d, t, q, m, dt, dd, and td”. These symbols indicate “singlet, doublet, triplet, quartet, multiplet, doublettriplet, doubletdoublet, and tripletdoublet”, respectively. All commercially obtained reagents were employed as received. Analytical TLC was carried out using pre-coated silica gel plates (Wako TLC Silicagel 70F254, FUJIFILM Wako Pure Chemical Corporation, Osaka, Japan). Wakogel 60N 63–212 μm was used for column chromatography.

2.2. Total Synthesis of Multifidene (1)

2.2.1. Alcohol S1

To a solution of cis-2-butane-1, 4-diol (1.85 g, 21.0 mmol) in DMF (30 mL) were added imidazole (4.28 g, 63.0 mmol) and TESCl (7.60 mL, 50.4 mmol) at 0 °C under an Ar atmosphere. The mixture was stirred at room temperature for 16 h, quenched with EtOH and saturated NaHCO3, extracted with AcOEt, washed with water and brine, dried over Na2SO4, filtered through short silica gel pad and concentrated in vacuo. The crude TES ether was employed directly in the next reaction.
Ozone was bubbled through a solution of the crude TES ether in MeOH-CH2Cl2 (40 mL, 19:1 v/v) at −78 °C for 4 h. To the mixture was added PPh3 (5.50 g, 21.0 mmol). The mixture was warmed to room temperature, stirred for 1 h, dried over Na2SO4 and concentrated in vacuo. The crude aldehyde 12 was employed directly in the next reaction.
To a solution of the crude aldehyde 12 in THF-tBuOH (80 mL, 1:1 v/v) were added MeNO2 (2.56 mL, 46.2 mmol) and KOtBu (471 mg, 8.40 mmol) at room temperature. The mixture was stirred for 1 h, quenched with saturated NaHCO3, extracted with AcOEt, dried over Na2SO4, filtered and concentrated in vacuo. The residue was purified by silica gel column chromatography (hexane: EtOAc = 99:1, 98:2, 96:4 then 94:6) to obtain the alcohol S1 (6.40 g, 27.2 mmol, 65% over 3 steps) as a colorless oil: IR (neat) 3446, 2956, 2913, 2878, 1556, 1458, 1416, 1379, 1240, 1206, 1122, 1006, 963, 943, 886, 804, 744, 674, 635 cm−1; 1H NMR (CDCl3, 400 MHz) δ 0.62 (6H, q, J = 7.8 Hz), 0.96 (9H, t, J = 7.8 Hz), 2.76 (1H, d, J = 6.8 Hz), 3.70 (2H, d, J = 4.9 Hz), 4.34–4.41 (1H, m), 4.45–4.56 (2H, m); 13C NMR (CDCl3, 100 MHz) δ 4.1, 6.6, 63.3, 68.9, 77.9; HRMS (ESI-orbitrap) m/z: [M+Na]+; Calcd for C9H21NO4NaSi 258.1132; Found 258.1136.

2.2.2. Adduct 7

To a solution of S1 (1.14 g, 4.84 mmol) in CH2Cl2 (16 mL) were added NEt3 (1.68 mL, 12.1 mmol) and MsCl (525 μL, 6.77 mmol) at 0 °C under Ar atmosphere. The mixture was stirred for 5 min, quenched with saturated NH4Cl, extracted with AcOEt, washed with brine, dried over Na2SO4, filtered through short silica gel pad and concentrated in vacuo. The crude olefin 7 was employed directly in the next reaction.
To a solution of the crude olefin 7 in PhMe (10 mL) were added 4-pentenal (9) (573 μL, 5.81 mmol), p-nitrophenol (134 mg, 0.968 mmol) and catalyst S-8 (157 mg, 0.484 mmol) in PhMe (3.0 mL) at −15 °C to room temperature under Ar atmosphere. The mixture was stirred for 22 h and purified by silica gel column chromatography (hexane: EtOAc = 99.5:0.5, 99:1 then 98:2) to obtain the adduct 7 (1.25 g, 4.15 mmol, 85% over 2 steps) as a colorless oil: [α]D26 + 32.9 (c 1.84, CHCl3); IR (neat) 2956, 2912, 2877, 2735, 1725, 1642, 1556, 1457, 1416, 1378, 1240, 1104, 1005, 920, 797, 745 cm−1; 1H NMR (CDCl3, 400 MHz) δ 0.59 (6H, q, J = 7.8 Hz), 0.94 (9H, t, J = 7.8 Hz), 2.30–2.37 (1H, m), 2.50–2.57 (1H, m), 2.67 (1H, q, J = 6.8 Hz), 2.81–2.86 (1H, m), 3.70 (2H, d, J = 5.4 Hz), 4.46 (1H, dd, J = 12.7, 4.9 Hz), 4.57 (1H, dd, J = 12.7, 8.3 Hz), 5.13 (1H, d, J = 9.8 Hz), 5.14 (1H, d, J = 17.1 Hz), 5.71–5.79 (1H, m); 13C NMR (CDCl3, 100 MHz) δ 4.1, 6.6, 30.4, 39.6, 50.1, 60.8, 74.1, 76.6, 77.0, 77.3, 118.3, 134.1, 202.1; HRMS (ESI-orbitrap) m/z: [M+Na]+; Calcd for C14H27NO4NaSi 324.1612; Found 324.1608.

2.2.3. Diolefin 13

To a solution of MePh3PBr (3.53 g, 9.90 mmol) in THF (10 mL) was added NaHMDS (1.0 M in THF, 9.00 mL, 9.00 mmol) at 0 °C under Ar. After 10 min, 7 (1.25 g, 4.15 mmol) in THF (10 mL) was added to the mixture at −40 °C. The mixture was stirred for 10 min, warmed to 0 °C, quenched with saturated NH4Cl, extracted with AcOEt, washed with brine, dried over Na2SO4, filtered and concentrated in vacuo. The residue was purified by silica gel column chromatography (hexane: EtOAc = 99:1 then 98:2) to obtain the diolefin 13 (1.10 g, 3.67 mmol, 88%) as a colorless oil: [α]D26 −12.9 (c 1.51, CHCl3); IR (neat) 3078, 2956, 2912, 2877, 1640, 1554, 1457, 1418, 1377, 1239, 1109, 1002, 918, 797, 744, 677 cm−1; 1H NMR (CDCl3, 400 MHz) δ 0.55 (6H, q, J = 7.8 Hz), 0.91 (9H, t, J = 7.8 Hz), 2.05–2.13 (1H, m), 2.21–2.28 (2H, m), 2.32–2.37 (1H, m), 3.64 (2H, d, J = 5.4 Hz), 4.34–4.43 (2H, m), 4.99–5.11 (4H, m), 5.50–5.59 (1H, m), 5.63–5.73 (1H, m); 13C NMR (CDCl3, 100 MHz) δ 4.2, 6.6, 35.9, 42.3, 42.4, 60.6, 75.2, 76.6, 77.0, 77.3, 116.8, 117.7, 135.6, 138.3; HRMS (ESI-orbitrap) m/z: [M+Na]+; Calcd for C15H29NO3NaSi 322.1813; Found 322.1809.

2.2.4. Triolefin 6

To a solution of 13 (164 mg, 0.548 mmol) in TFE-DCE (3.0 mL, 9:1 v/v) were added fluorescein (18.2 mg, 0.0548 mmol) and Cs2CO3 (357 mg, 1.09 mmol) at −20 °C under O2 atmosphere. The mixture was stirred for 46 h under visible light irradiation (ν max = 470 nm, 610 mW), filtered through short path silica gel pad and concentrated in vacuo to obtain the crude aldehyde 14, which was employed directly in the next reaction.
To a solution of MePh3PBr (360 mg, 1.03 mmol) in THF (5.0 mL) was added NaHMDS (1.0 M in THF, 1.00 mL, 1.00 mmol) at 0 °C under Ar. After 5 min, the crude aldehyde 14 in THF (2.5 mL) was added to the mixture at −40 °C. The mixture was stirred for 10 min at −40 °C and 20 min at 0 °C, quenched with saturated NH4Cl, extracted with AcOEt, washed with brine, dried over Na2SO4, filtered and concentrated in vacuo. The residue was purified by silica gel column chromatography (hexane: EtOAc = 99.8:0.2 then 99.5:0.5) to obtain the triolefin 6 (70.8 mg, 0.266 mmol, 48% over 2 steps) with recovery of 13 (15.2 mg, 9%): a colorless oil; [α]D26 −35.6 (c 0.59, CHCl3); IR (neat) 3076, 2955, 2911, 2877, 1640, 1457, 1416, 1380, 1239, 1097, 994, 913, 803, 742, 669 cm−1; 1H NMR (CDCl3, 400 MHz) δ 0.58 (6H, q, J = 7.8 Hz), 0.95 (9H, t, J = 7.8 Hz), 2.10–2.16 (2H, m), 2.29–2.35 (1H, m), 2.40–2.47 (1H, m), 3.49–3.58 (2H, m), 4.97–5.10 (6H, m), 5.53–5.64 (2H, m), 5.69–5.77 (1H, m); 13C NMR (CDCl3, 100 MHz) δ 4.4, 6.8, 37.1, 43.2, 49.6, 64.2, 115.6, 116.2, 117.4, 136.4, 137.3, 138.7; HRMS (ESI-orbitrap) m/z: [M+Na]+; Calcd for C16H30ONaSi 289.1964; Found 289.1956.

2.2.5. Cyclopentene 15

To a solution of 6 (22.3 mg, 0.0837 mmol) in CH2Cl2 (3.0 mL) was added Grubbs 1st generation catalyst (3.4 mg, 0.00413 mmol) at room temperature under Ar atmosphere. After 15 h, MeOH (33.4 μL, 0.837 mmol) and CSA (3.80 mg, 0.0165 mmol) were added. The mixture was further stirred for 90 min, quenched with Et3N (2.30 μL, 0.0165 mmol) and concentrated in vacuo. The crude product was purified by silica gel column chromatography (pentane: Et2O = 95:5 then 90:10) to obtain the cyclopentene 15 (9.90 mg, 0.0797 mmol, 95%) as a colorless oil: [α]D23 + 180 (c 0.20, CHCl3); IR (neat) 3361, 2923, 2853, 1540, 1507, 1024, 910, 669 cm−1; 1H NMR (CDCl3, 400 MHz) δ 2.24–2.31 (1H, m), 2.48 (1H, dd, J = 16.6, 8.3 Hz), 2.87–2.94 (1H, m), 3.03 (1H, quin, J = 8.3 Hz), 3.60 (2H, d, J = 5.4 Hz), 5.07 (1H, dd, J = 10.2, 1.9 Hz), 5.14 (1H, d, J = 17.1 Hz), 5.64–5.65 (1H, m), 5.88–5.91 (1H, m), 6.03 (1H, dt, J = 17.1, 8.8 Hz); 13C NMR (CDCl3, 100 MHz) δ 38.2, 44.6, 51.4, 62.8, 115.5, 130.8, 132.5, 139.6; HRMS (ESI-orbitrap) m/z: [M+Na]+; Calcd for C8H12ONa 147.0786; Found 147.0782.

2.2.6. Multifidene (1)

To a solution of 15 (46.7 mg, 0.376 mmol) in CH2Cl2 (1.8 mL) was added DMP (191 mg, 0.450 mmol) at 0 °C. The reaction mixture was stirred for 25 min, quenched with saturated NaHCO3: 20% Na2S2O3 (1:1, v/v) and extracted with Et2O. The combined extracts were washed with brine, dried over Na2SO4, filtered, and concentrated in vacuo to obtain the crude aldehyde, which was employed directly in the next reaction.
To a solution of nPrPh3PBr (385 mg, 1.00 mmol) in Et2O (3.0 mL) was added NaHMDS (1.0 M in THF, 0.900 mL, 0.900 mmol) at 0 °C under Ar. After 10 min, the crude aldehyde in Et2O (2.0 mL) was added to the mixture at −78 °C. The mixture was stirred for 10 min at −78 °C and 20 min at 0 °C, quenched with saturated NH4Cl, extracted with Et2O, washed with brine, dried over Na2SO4, filtered and concentrated in vacuo. The residue was purified by silica gel column chromatography (pentane only) to obtain multifidene 1 (15.9 mg, 0.106 mmol, 28% over 2 steps) as a colorless oil: [α]D23 + 162.5 (c 0.20, pentane); 1H NMR (CDCl3, 400 MHz) δ 0.96 (3H, t, J = 7.8 Hz), 2.03–2.10 (2H, m), 2.24–2.31 (1H, m), 2.41–2.51 (1H, m), 2.97 (1H, quint., J = 7.8 Hz), 3.59–3.64 (1H, m), 4.91–5.02 (2H, m), 5.13 (1H, tt, J = 11.0, 1.5 Hz), 5.39 (1H, dt, J = 11.0, 7.3 Hz), 5.58–5.61 (1H, m), 5.76–5.78 (1H, m), 5.87 (1H, ddd, J = 17.3, 10.3, 8.1 Hz); 13C NMR (CDCl3, 100 MHz) δ 14.4, 20.8, 37.1, 46.7, 46.8, 114.0, 128.3, 129.4, 131.9, 134.3, 140.2.

3. Results and Discussion

Our synthetic aim is to develop a flexible synthetic route to multifidene (1) in order to access derivatives such as the enantiomer of 1 or a molecular probe to investigate the molecular mechanism of the pheromone functionality. The retrosynthetic analysis toward 1 is shown in Scheme 1. It was envisioned that asymmetric conjugate addition between aldehyde 6 and nitroolefin 7 by the Hayashi–Jørgensen catalyst S-8 [25] would construct adduct 9 with the contiguous stereoconfiguration of 1. After the conversion of the aldehyde and nitro groups into the corresponding double bonds in triolefin 10, a ring-closing metathesis reaction (RCM) and subsequent Wittig reaction provides 1.
The synthesis started with the commercially available cis-2-butene-1, 4-diol (11) (Scheme 2). The protection of the two hydroxy groups of 11 with TES groups and ozonolysis of the double bond afforded aldehyde 12. The treatment of 12 with MeNO2 and KOtBu furnished an alcohol, which was dehydrated to provide nitroolefin 10. The subsequent addition reaction with commercially available 4-pentenal (6) in the presence of the Hayashi–Jørgensen catalyst S-8 at −15 °C produced adduct 7 with high stereoselectivity (>95% ee and 7:1 dr) and high yield (85% over 2 steps) [25,26] (see Supplementary Materials for the enantiopurity of 7). The newly formed absolute configurations were determined by final conversion to 1, and diastereoselectivity was estimated by 1H NMR analysis. In this reaction, the lower temperature was important because lower diastereoselectivity was observed in the 1H NMR spectrum at room temperature or 0 °C. The aldehyde in 7 was transformed into a vinyl group through a Wittig reaction.
With the diolefin 13 in hand, the Nef reaction, i.e., conversion of the nitro group to an aldehyde, was investigated next. The previously reported conditions, such as NaOMe-H2SO4 [27], DABCO-O2 [28], KMnO4-Base (K2CO3 [29] or KOH [30]), TiCl3 [31] or Cu (0)-TMEDA-O2 [32], resulted in undesired products or low yields. In the pioneering publication on the Nef reaction in 1986 to form aldehydes, rose bengal enabled the conversion of a nitro group to an aldehyde in the presence of molecular oxygen and irradiation with a tungsten lamp [33]. Based on the recent progress on photoreactions in this decade, we attempted to optimize this reaction. The results of the investigation are summarized in Table 1. First, bases (NaOMe, KOtBu, and Cs2CO3) were investigated (entries 1–6). The reactions were carried out with 1.5 equivalents of base (NaOMe, KOtBu, or Cs2CO3) in MeOH-DCE (1,2-dichloroethane) or TFE (2,2,2-trifluoroethanol)-DCE solvent in the presence of a catalytic amount of fluorescein under visible light irradiation (470 nm, 620 mW) for 8 h at 0 °C. The addition of the less polar solvent, DCE, was essential to dissolve 13 in an alcoholic solvent. The Nef reactions with NaOMe were less effective than other bases, producing the desired aldehyde 14 as a minor component (entries 1, 2). When KOtBu was used in MeOH-DCE, starting material 13 was completely consumed, although a considerable amount of side products was observed in the 1H NMR spectrum of the crude product (entry 3). The structures of the side products could not be determined due to the complexity of the mixture. The reaction in TFE-DCE produced a mixture of 13 and 14 in a ratio of 45:55 along with trace amounts of side products (entry 4). Cs2CO3 showed a similar tendency as KOtBu, but the ratio was improved to 33:67 in TFE-DCE solvent (entries 5, 6). The reaction in EtOH with Cs2CO3 (entry 7) was similar to that in MeOH. Other solvents, iPrOH or MeCN, resulted in a lower conversion of 13 (entries 8, 9).
A possible reaction pathway based on the previous work is described in Scheme 3 [33]. The abstraction of an α-hydrogen of the nitro group produces nitronate A. The subsequent reaction of A with singlet molecular oxygen generated from O2 and fluorescein under light irradiation produces adduct B through [2+2] cycloaddition. Adduct B dissociates to furnish an aldehyde. The above results indicated that effects of the solvent were crucial for the conversion of 13. A primary alcohol efficiently promoted the Nef reaction when Cs2CO3 was employed as a base. It is assumed that a less bulky base generated from the alcohol and Cs2CO3 can readily abstract an α-hydrogen. Among the reactions with Cs2CO3, those in EtOH and TFE were noteworthy. As shown in entries 6 and 7, EtOH completely consumed 13 within 8 h, although side products were formed, whereas TFE exhibited a slower reaction rate than EtOH. We believe this slower reaction inhibits the formation of side products, and generates a less basic and/or less nucleophilic profile [34] of 2, 2, 2-trifluoroethoxide, which affects the reaction rate.
The reaction was further examined for the efficient formation of 14 by screening equivalents of Cs2CO3 and reaction temperatures with the TFE-DCE solvent system. Two equivalents of Cs2CO3 consumed the starting material; however, the partial epimerization of the product at the α-position was observed (entry 10). Although a longer reaction time (16 h) was required, a lower temperature (−10 °C) produced a similar result as that at 0 °C (entry 11). The Nef reaction at −20 °C for 16 h provided a 44:56 mixture of 13 and 14 (entry 12), and the ratio of 14 at −20 °C was improved through a much longer reaction time (48 h) (entry 13). The efficiency of this Nef reaction was affected by reaction scale; however, for this entry, more than 100 mg of 13 was used for the effective total synthesis of 1.
With the optimized conditions for the Nef reaction in hand, we completed the total synthesis of multifidene (1), as shown in Scheme 4. The resultant labile aldehyde 14 was transformed into triolefin 6 in 48% yield over two steps from 13. The ring-closing metathesis reaction with the Grubbs 1st generation catalyst [35] proceeded regioselectively to produce the desired cyclopentene 15 after the removal of the TES group under mild acidic conditions [36]. We rationalized this regioselectivity by kinetic control. The desired product is the favored five-membered ring, but other possible products are much less favored, such as a four-membered ring. Cyclopentene 15 was successfully transformed into multifidene (1) in 28% yield over two steps: Dess–Martin oxidation and Z-selective Wittig reaction. In the final steps, we assumed that a low yield is due to a partial isomerization reaction of the labile β,γ-unsaturated aldehyde to the α,β-unsaturated aldehyde under the strong basic conditions of the Wittig reaction because a spot having UV absorption was observed in TLC analysis in the Wittig reaction. This spot was not found in the oxidation reaction. Although a one-pot operation of Swern oxidation and Wittig reaction as non-isomerization conditions have been reported [16], this operation in our study resulted in the recovery of the aldehyde. Some bases were also screened to result in a lower yield or complex mixture. 1H and 13C NMR spectra of synthetic 1 were identical with those of the previously prepared compound [17].

4. Conclusions

In summary, the enantioselective total synthesis of multifidene was described. The conjugate addition of an aldehyde to a nitroolefin in the presence of a Hayashi–Jørgensen catalyst and the newly developed Nef reaction producing a labile aldehyde with fluorescein under visible light irradiation enabled the efficient synthesis. The developed synthetic route opens up an opportunity to prepare the enantiomer or diastereomers of multifidene, or other derivatives with a cyclohexene core or a molecular probe. We are currently investigating a bioassay to determine the relationship between chemotaxis and phototaxis of this pheromone with brown algae.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/org3030015/s1: determination of enantiopurity of adduct 9, and the NMR spectra of the synthetic samples.

Author Contributions

T.U. and N.K.-T. designed the study; T.U. and M.H. synthesized the compounds; T.U. acquired the funds; T.U., N.K.-T. and F.M. analyzed the data and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by The Akiyama Life Science Foundation (2019-5).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jaenicke, L.; Müller, D.G.; Moore, R.E. Multifidene and Aucantene, C11 Hydrocarbons in the Male-Attracting Essential Oil from the Gynogametes of Cutleria multifida (Phaeophyta). J. Am. Chem. Soc. 1974, 96, 3324–3325. [Google Scholar]
  2. Müller, D.G.; Jaenicke, L.; Donike, M.; Akintobi, T. Sex Attractant in a Brown Alga: Chemical Structure. Science 1971, 171, 815–817. [Google Scholar]
  3. Moore, R.E.; Pettus, J.A., Jr. Isolation and Structure Determination of Dictyopterenes C′ and D′ from Dictyopteris. Stereospecificity in the Cope Rearrangement of Dictyopterenes A and B. J. Am. Chem. Soc. 1971, 93, 3087–3088. [Google Scholar]
  4. Moore, R.E.; Pettus, J.A., Jr.; Doty, M.S. Dictyopterene, A. An Odoriferous Constituent from Algae of the Genus Dictyopteris. Tetrahedron Lett. 1968, 9, 4787–4790. [Google Scholar]
  5. Pettus, J.A., Jr.; Moore, R.E. Isolation and Structure Determination of an Undeca-1,3,5,8-tetraene and Dictyopterene B from Algae of the Genus Dictyopteris. J. Chem. Soc. D. 1970, 1970, 1093–1094. [Google Scholar]
  6. Jaenicke, L.; Boland, W. Signal Substances and Their Reception in the Sexual Cycle of Marine Brown Algae. Angew. Chem. Int. Ed. Engl. 1982, 94, 643–653. [Google Scholar]
  7. Maier, I.; Müller, D.G. Sexual Pheromones in Algae. Bio. Bull. 1986, 170, 145–175. [Google Scholar]
  8. Boland, W. Chemische Kommunikation bei der Sexuellen Fortpflanzung Mariner Braunalgen. Biologie in Unserer Zeit 1987, 17, 176–185. [Google Scholar]
  9. Jaenicke, L. One Hundred and One Years of Chemotaxis. Pfeffer, Pheromones, and Fertilization. Bot. Acta 1988, 101, 149–159. [Google Scholar]
  10. Maier, I. Gamete Orientation and Induction of Gametogenesis by Pheromones in Algae and Plants. Plant Cell Environ. 1993, 16, 891–907. [Google Scholar]
  11. Derenbach, J.B.; Pesandeo, D. Investigations into a Small Fraction of Volatile Hydrocarbons: III. Two Diatom Cultures Produce Ectocarpene, a Pheromone of Brown Algae. Mar. Chem. 1986, 19, 337–432. [Google Scholar]
  12. Boland, W. The Chemistry of Gamete Attraction: Chemical Structures, Biosynthesis, and (A)biotic Degradation of Algal Pheromones. Proc. Natl. Acad. Sci. USA 1995, 92, 37–43. [Google Scholar]
  13. Jüttner, F.; Wurster, K. Evidence of Ectocarpene and Dictyopterenes A and C′ in the Water of a Freshwater Lake. Limnol. Oceanogr. 1984, 29, 1322–1324. [Google Scholar]
  14. Boland, W.; Marner, F.-J.; Jaenicke, L.; Müller, D.G.; Folster, E. Comparative Receptor Study in Gamete Chemotaxis of the Seaweeds Ectocarpus siliculosus and Cutleria multifida. Eur. J. Biochem. 1983, 134, 97–103. [Google Scholar]
  15. Hay, M.E.; Duffy, J.E.; Fenical, W.; Gustafson, K. Chemical Defense in the Seaweed Dictyopteris delicatula: Differential Effects against Reef Fishes and Amphipods. Mar. Ecol. Prog. Ser. 1988, 48, 185–192. [Google Scholar]
  16. Lebreton, J.; Alphand, V.; Furstoss, R. A Short Chemoenzymatic Synthesis of (+)-Multifidene and (+)-Viridiene. Tetrahedron Lett. 1996, 37, 1011–1014. [Google Scholar]
  17. Kramp, P.; Helmchen, G.; Holmes, A.B. Syntheses of Enantiomerically Pure ent-Multifidene and Related Compounds. J. Chem. Soc. Chem. Commun. 1993, 551–552. [Google Scholar]
  18. Wirth, D.; Fischer-Lui, I.; Boland, W.; Icheln, D.; Runge, T.; Konig, W.A.; Phillips, J.; Clayton, M. Unusual and Novel C11H16 Hydrocarbons from the Southern Australian Brown Alga Dictyopteris acuostichoides (Phaeophyceae). Helv. Chim. Acta 1992, 75, 734–744. [Google Scholar]
  19. Crouse, G.D.; Paquette, L.A. Total Synthesis of (±)-Multifidene, the Gamete Attractant of the Phaeophyte Cutleria multifida. J. Org. Chem. 1981, 46, 4272–4274. [Google Scholar]
  20. Boland, W.; Jaenicke, L.; Müller, D.G. Synthese und biologische Aktivitaten von (+)- und (−)-Multifiden. Liebigs Ann. Chem. 1981, 1981, 2266–2271. [Google Scholar]
  21. Boland, W.; Jaenicke, L. Stereoselective Synthesis of “Multifidenes” and Related Structures. Contribution to the Synthetic Chemistry of Cis-Disubstituted Alken-l-ylcyclopentenes. J. Org. Chem. 1979, 44, 4819–4824. [Google Scholar]
  22. Hemamalini, S.; Scheffold, R. Synthesis of (+)-Multifidene. Helv. Chim. Acta 1995, 78, 447–451. [Google Scholar]
  23. Kinoshita-Terauchi, N.; Shiba, K.; Umezawa, T.; Matsuda, F.; Motomura, T.; Inaba, K. A Brown Algal Sex Pheromone Reverses the Sign of Phototaxis by cAMP/Ca2+-Dependent Signaling in the Male Gametes of Mutimo cylindricus (Cutleriaceae). J. Photochem. Photobiol. B 2019, 192, 113–123. [Google Scholar]
  24. Boland, W.; Mertes, K. Biosynthesis of Algal Pheromones. Eur. J. Biochem. 1985, 147, 83–91. [Google Scholar]
  25. Hayashi, Y.; Gotoh, H.; Hayashi, T.; Shoji, M. Diphenylprolinol Silyl Ethers as Efficient Organocatalysts for the Asymmetric Michael Reaction of Aldehydes and Nitroalkenes. Angew. Chem. Int. Ed. 2005, 44, 4212–4215. [Google Scholar]
  26. Patora-Komisarska, K.; Benohoud, M.; Ishikawa, H.; Seebach, D.; Hayashi, Y. Organocatalyzed Michael Addition of Aldehydes to Nitro Alkenes—Generally Accepted Mechanism Revisited and Revised. Helv. Chim. Acta 2011, 94, 719–745. [Google Scholar]
  27. Humphrey, J.M.; Arnold, E.P.; Chappie, T.A.; Feltenberger, J.B.; Nagel, A.; Simon, W.; Suarez-Contreras, M.; Tom, N.J.; O’Neill, B.T. Diastereoselective Synthesis of 2,3,6-Trisubstituted Piperidines. J. Org. Chem. 2009, 74, 4525–4536. [Google Scholar]
  28. Umemiya, S.; Nishino, K.; Sato, I.; Hayashi, Y. Nef Reaction with Molecular Oxygen in the Absence of Metal Additives, and Mechanistic Insights. Chem. Eur. J. 2014, 20, 15753–15759. [Google Scholar]
  29. Vichare, P.; Chattopadhyay, A. Nitrolaldol Reaction of (R)-2,3-Cyclohexylideneglyceraldehyde: A Simple and Stereoselective Synthesis of the Cytotoxic Pachastrissamine (Jaspine B). Tetrahedron Asymmetry 2010, 21, 1983–1987. [Google Scholar]
  30. Muratake, H.; Natsume, N.; Nakai, H. Palladium-Catalyzed Intramolecular α-Arylation of Aliphatic Ketone, Formyl, and Nitro Groups. Tetrahedron 2004, 60, 11783–11803. [Google Scholar]
  31. Ghorai, S.K.; Hazra, N.K.; Mal, D. Facile Synthesis of 4-Functionalized Cyclopentenones. Synth. Commun. 2007, 37, 1949–1956. [Google Scholar]
  32. Balogh-Hergovich, É.; Kaizer, J.; Speier, G. Copper Mediated Conversion of Nitro Compounds to Aldehydes or Ketones by Dioxygen. Chem. Lett. 1996, 25, 573–574. [Google Scholar]
  33. Yamashita, M.; Nomoto, H.; Imoto, H. Preparation of Dimethyl (1-Formylalkyl)phosphonates via Singlet Oxygen Adducts. Synthesis 1987, 716–718. [Google Scholar]
  34. Minegishi, S.; Kobayashi, S.; Mayr, H. Solvent Nucleophilicity. J. Am. Chem. Soc. 2004, 126, 5174–5181. [Google Scholar]
  35. Schwab, P.; France, M.B.; Ziller, J.W.; Grubbs, R.H. A Series of Well-Defined Metathesis Catalysts-Synthesis of RuCl2(=CHR′)-(PR3)2 and Its Reactions. Angew. Chem. Int. Ed. Engl. 1995, 34, 2039–2041. [Google Scholar]
  36. Umezawa, T.; Mizutani, N.; Matsuo, K.; Tokunaga, Y.; Matsuda, F.; Nehira, T. Assignment of Absolute Configuration of Bromoallenes by Vacuum-Ultraviolet Circular Dichroism (VUVCD). Molecules 2021, 26, 1296. [Google Scholar]
Figure 1. Chemical structures of C11 pheromones.
Figure 1. Chemical structures of C11 pheromones.
Organics 03 00015 g001
Figure 2. Previous syntheses of optically active multifidene.
Figure 2. Previous syntheses of optically active multifidene.
Organics 03 00015 g002
Figure 3. Proposed biosynthesis of C11 pheromones.
Figure 3. Proposed biosynthesis of C11 pheromones.
Organics 03 00015 g003
Scheme 1. Retrosynthetic analysis.
Scheme 1. Retrosynthetic analysis.
Organics 03 00015 sch001
Scheme 2. Synthesis of the optically active olefin 13.
Scheme 2. Synthesis of the optically active olefin 13.
Organics 03 00015 sch002
Scheme 3. Possible reaction mechanism.
Scheme 3. Possible reaction mechanism.
Organics 03 00015 sch003
Scheme 4. Total synthesis of multifidene.
Scheme 4. Total synthesis of multifidene.
Organics 03 00015 sch004
Table 1. Optimization of the Nef reaction 1.
Table 1. Optimization of the Nef reaction 1.
Organics 03 00015 i001
EntryBase (Equiv.)SolventTimeTemp.Ratio (13:14) 2
1NaOMe (1.5)MeOH-DCE (9:1)8 h0 °C76:24
2NaOMe (1.5)TFE-DCE (9:1)8 h0 °C91:9
3KOtBu (1.5)MeOH-DCE (9:1)8 h0 °C0:100 3
4KOtBu (1.5)TFE-DCE (9:1)8 h0 °C45:55
5Cs2CO3 (1.5)MeOH-DCE (9:1)8 h0 °C0:100 3
6Cs2CO3 (1.5)TFE-DCE (9:1)8 h0 °C33:67
7Cs2CO3 (1.5)EtOH-DCE (9:1)8 h0 °C0:100 3
8Cs2CO3 (1.5)iPrOH-DCE (9:1)8 h0 °C86:14
9Cs2CO3 (1.5)MeCN-DCE (9:1)8 h0 °C96:4
10Cs2CO3 (2.0)TFE-DCE (9:1)8 h0 °C0:100 4
11Cs2CO3 (2.0)TFE-DCE (9:1)16 h−10 °C0:100 4
12Cs2CO3 (2.0)TFE-DCE (9:1)16 h−20 °C44:56
13Cs2CO3 (2.0)TFE-DCE (9:1)48 h−20 °C13:87 5
1 40–50 mg (ca 0.14 mmol) of 13 was used for the optimization, otherwise mentioned; 2 The ratio of 13 and 14 was estimated with crude 1H NMR; 3 Considerable amount of side products were observed; 4 Partial epimerization was observed; 5 164 mg of 13 was used.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Umezawa, T.; Hara, M.; Kinoshita-Terauchi, N.; Matsuda, F. Enantioselective Total Synthesis of Multifidene, a Sex Pheromone of Brown Algae. Organics 2022, 3, 187-195. https://doi.org/10.3390/org3030015

AMA Style

Umezawa T, Hara M, Kinoshita-Terauchi N, Matsuda F. Enantioselective Total Synthesis of Multifidene, a Sex Pheromone of Brown Algae. Organics. 2022; 3(3):187-195. https://doi.org/10.3390/org3030015

Chicago/Turabian Style

Umezawa, Taiki, Misaki Hara, Nana Kinoshita-Terauchi, and Fuyuhiko Matsuda. 2022. "Enantioselective Total Synthesis of Multifidene, a Sex Pheromone of Brown Algae" Organics 3, no. 3: 187-195. https://doi.org/10.3390/org3030015

Article Metrics

Back to TopTop